You are on page 1of 9

Rheol Acta (2008) 47:425–433

DOI 10.1007/s00397-008-0265-4

ORIGINAL CONTRIBUTION

Rheology of carbon nanofiber-reinforced polypropylene


Simona Ceccia & Dino Ferri & Daniela Tabuani &
Pier Luca Maffettone

Received: 5 June 2007 / Revised: 21 December 2007 / Accepted: 28 January 2008 / Published online: 15 April 2008
# Springer-Verlag 2008

Abstract The rheological properties of two different nano- mechanical, thermal, and electrical properties (e.g., Yang
composite systems consisting in the dispersion of carbon et al. 2007; Kelarakis et al. 2006; Gao et al. 2006; Choi et
nanofibers (CNFs) in polypropylene are investigated. The al. 2005; Xu et al. 2004). CNFs have multiple concentric
nanoreinforced systems were identically prepared with two walls, with diameter ranging from 70 to 200 nm and length
CNFs that differ only in the length of the fibers being 50–100 μm. They are not continuous tubes as the graphene
otherwise identical to analyze the effect of fiber aspect cylinders, and they are not parallel to the fiber axis. Rather,
ratio. Linear dynamic viscoelasticity and the steady-state they show a 20° angle with respect to fiber axis in a Dixie
rheology of the two different nanocomposites are presented. cup arrangement terminating at the wall of the next outer
The system reinforced with CNFs with larger aspect ratio tube (Kang et al. 2006). The characteristic nanoscopic
shows several rheological features that resemble peculiar- dimension together with a relatively low cost and the easy
ities of rodlike polymers in the nematic liquid crystalline incorporation into polymers make CNFs an obvious
phase. candidate for the production of high performance light
materials. Several recent papers show the benefits of the
Keywords Carbon nanofiber . Viscosity . Normal stresses . addition of CNFs in polymer matrices in terms of
Linear viscoelasticity mechanical, electrical, and thermal properties. For example,
Kumar et al. (2002) showed that fibers from PP/CNF
composites can be spun using conventional equipment, and
Introduction possess superior modulus and compressive strength with
respect to the pure polymer at only 5 wt% of CNFs. Such
Carbon nanofibers (CNFs) represent a viable solution to the an improvement was obtained by dispersing as received
preparation of polymer nanocomposites with improved CNFs into the polymer via melt processing. Sandler et al.
(2003) studied nanoreinforced fibers of a semicrystalline
high-performance poly(etheretherketone) containing up to
S. Ceccia : D. Tabuani 10 wt% CNFs. The carbon nanofibers were found to be
Centro Ingegneria Materie Plastiche,
Viale T. Michel 5,
well aligned with the direction of flow during processing,
15100, Alessandria, Italy and, correspondingly, nanocomposite stiffness, yield stress,
and fracture strength improved with respect to neat
D. Ferri polymer. Upon addition of nanofibers, a significant increase
Centro Ricerche “Claudio Buonerba”, Polimeri Europa,
Via Taliercio 14,
of the degree of crystallinity of the matrix was also
46100, Mantova, Italy observed. Gauthier et al. (2005) studied the reinforcement
of rubbery matrices by CNFs and showed that the
P. L. Maffettone (*) mechanical performances revealed a linear increase of the
Dipartimento di Ingegneria Chimica Università Federico II Napoli,
Piazzale V. Tecchio 80,
modulus measured above and below the glass transition
80125, Napoli, Italy temperature for nanofiber content up to 10 wt%. In the case
e-mail: pierluca.maffettone@unina.it of an epoxy matrix, the ultimate stress and strain were also
426 Rheol Acta (2008) 47:425–433

largely increased, even for very low-fiber content (1 wt%). Also, the experimental results were well described by a
Shen et al. (2006) showed that CNFs can be used both as model similar to those commonly adopted to describe fiber
nucleant and reinforcement for polystyrene (PS) foams. The suspensions in viscoelastic liquids.
inclusion of CNFs exhibited a substantial impact on the In the present work, we characterize the rheological
morphology and properties of PS foams by decreasing response of two different nanocomposite systems consisting
the cell size and increasing the cell density. A substantial in the dispersion of CNFs in polypropylene. The nano-
strength enhancement of PS foams due to the incorporation reinforced systems were obtained with two CNFs that differ
of CNFs was experimentally observed. only for the length of the fibers being otherwise identical
The rather common observation of improvements of the and identically prepared. In such a way, a clear analysis of
mechanical properties of CNF-reinforced plastic products is the relevance of the fiber aspect ratio can be addressed. The
generally related to the formation of a microstructure. In the linear dynamic viscoelasticity and the steady state rheology
molten state, such microstructure can be easily probed with of the two different nanocomposites are presented. As it
rheological experiments. Linear and nonlinear viscoelastic will be shown, the rheological characterization of system
properties provide important information on the dispersion reinforced with CNFs with larger aspect ratio shows several
of the filler in the matrix, on the concentration and size features that resemble peculiarities of rodlike polymers in
distribution of the filler, and its wettability. Only few and the nematic liquid crystalline phase. This phenomenology is
very recent papers, however, were devoted to the rheolog- similar to that recently observed for carbon nanotubes,
ical characterization of CNF-reinforced polymers (Lozano which are considered as a type of highly conjugated, rigid-
et al. 2004; Xu et al. 2005; Wang et al. 2006). Lozano et al. rod macromolecules. Carbon nanotubes have shown inter-
(2004) characterized the linear dynamic viscoelasticity of esting similarities with “rigid-rod” macromolecular systems
CNF nanoreinforced HDPE. They noted a monotonic (Song and Windle 2005). Indeed, it was reported that
increase of G’ and G” moduli both with frequency and carbon nanotubes can form nematic liquid crystalline
concentration, with the appearance of a solid-like plateau at phases (Song et al. 2003; Davis et al. 2004; Song and
low frequencies at loadings ≥20 wt%. Xu et al. (2005) Windle 2005). The existence of a nematic phase could of
presented a very detailed study on samples of CNFs in course lead to interesting final properties as large-scale
glycerol/water to analyze the effects of different purification alignment could be achieved under processing conditions.
techniques (sonication and acid treatment). Nanocompo-
sites from as-received nanofibers showed large aggregates
(mm to cm), whereas sonication significantly reduced Materials and methods
aggregate dimensions (50 μm), and an acid treatment
improved the dispersion but at the cost of shortening the The nanoreinforced polymer was prepared with a homo-
CNFs. A solid-like behavior of the moduli was found for polymer commonly used to produce injection-molded
loadings larger than 3 wt%. Xu et al. (2005) also modeled products (Moplen HP400R, Basell). The relevant polymer
the nanocomposite systems with either elastic or rigid properties are reported in Table 1.
dumbbells in a Newtonian solvent with isotropic or Two different nanofibers produced by Applied Sciences
anisotropic hydrodynamic drag, with or without hydrody- Inc were dispersed in the matrix: PR-24 LHT HD (High
namic interaction. They found that the elastic model with Density) and PR-24 LHT LD (Low Density). The HD and
anisotropic hydrodynamic drag and negligible hydrody- LD designations do not refer to the density of the individual
namic interaction successfully captures the rheological fiber but to the bulk density of the carbon nanofibers. The
behavior observed in their experiments. Very recently, the bulk, or aggregate density, of the nanofiber is controlled by
same research group (Wang et al. 2006) presented an altering the mixing intensity and duration of the debulking
analysis on shear rheology of nanofiber/polystyrene com- process. The debulking process reduces the length of the
posites. The focus was on the effect of the preparation individual fibers, with high density fiber having shorter
technique as captured by rheological measurements. It was fiber lengths than the low density fiber. The producer states
shown that melt blending strongly reduces the length of the
fibers with respect to solvent-casting preparation. A general Table 1 Properties of the polymeric matrix
monotonic increase of both G’ and G” was observed. The Property Method Value Unit
melt phase of solvent-cast composites with higher CNF
concentrations exhibits a plateau of the elastic modulus at Density (23 °C) ISO 1183 0.905 g/cm3
low frequencies, an apparent yield stress, and large first Melt flow rate ISO 1133 25 g/10 min
(230 °C/2.16 kg)
normal stress difference, N1, at low strain rates. The authors
Melt volume flow rate ISO 1133 34 cm3/10 min
ascribed such features to contact-based network nano-
(230 °C/2.16 kg)
structure which forms in the presence of longer CNFs.
Rheol Acta (2008) 47:425–433 427

Table 2 Relevant properties of the CNFs

Name N2 surface area, Moisture Iron content PAH content Diameter Length Nominal aspect
(m2/gm) (%) (PPM) (mg PAH/g fiber) (nm) (μm) ratio

PR-24 LHT HD 35–45 <5 <14,000 <1 100–200 50–100 250–1,000


PR-24 LHT LD 35–45 <5 <14,000 <1 100–200 100–200 500–2000

that typically, the fiber length in the low density product is chanical degradation. Thus, for the sake of comparison, the
100–200 μm, while the fiber length in the high density pure polypropylene matrix MPHP400R was also subjected
product is 50–100 μm (Burton, personal communication). to the same thermal and mechanical treatments as the
The relevant parameters of the nanofibers are reported in nanocomposites to have a consistent reference matrix.
Table 2. The samples were rheologically characterized with
The dispersion of nanofibers in polypropylene was different instruments. The linear viscoelastic storage moduli
achieved by melt compounding as-received CNFs with the were measured in shear at 25 °C with a Rheometrics
polymeric matrix in a Brabender internal mixer. The Mechanical Spectrometer (RMS) model 800 at a frequency
internal mixer consisted of a heated chamber with a volume of 1 Hz using the torsion rectangular geometry. The
of 55 cm3. The chamber was closed between two plates: a instrument was equipped with a force rebalance transducer
fixed plate with two counter-rotating mini-screws (maxi- able to detect torques within the range 0.02–20 N. The
mum rate 200 rpm), and a mobile one with the hopper and actuator was an air bearing motor with strain amplitude of
the chamber. Mixing was performed at 180 °C, 60 rpm, for 0.05–500 mrad and a rotation angular frequency varying
10 min. Processing conditions were chosen to optimize between 10–3 and 102 rad/s. The specimens were compres-
nanofiber dispersion and to minimize matrix degradation. sion-molded at about 210 °C and recovered as rectangular
Both nanocomposite systems were prepared following the 60×2×8 mm stripes to fit the RMS tools. The temperature
very same procedure. The melt compounding preparation is was stable within 0.2 °C over the range used in this study.
expected to cause the breakage of nanofibers, thus altering Strain sweeps were performed with strain amplitudes
their nominal properties. Transmission electron microscopy ranging from 0.01 to 1.4. The shear storage moduli reported
images recorded on CNF extracted by dissolving PP in in the following are those measured in the linear viscoelas-
Xylene (not shown here) showed that the ratio between tic response region. All tests were performed in N2
average LD-CNF and HD-CNF length is practically atmosphere.
preserved after preparation. For dynamic linear viscoelasticity of the melt, the
Five different compositions were prepared with HD samples were compression-molded at 210 °C under a
nanofibers as reported in Table 3; nine compositions were pressure of 20 MPa and recovered as 25 mm diameter
prepared with LD CNFs as reported in Table 4. disks to fit the RMS tools. The linear viscoelastic functions
Scanning electron microscopy (SEM) imaging was were measured in the RMS using the parallel plate
obtained by means of a LEO 1450 VP instrument on cryogenic geometry (diameter 25 mm and gaps between 0.6 and 1
fracture surfaces. SEM micrographs showed that a rather mm). Isothermal frequency scans were performed in the
uniform dispersion was generally obtained, but some aggre- range between 10−1 and 102 rad/s. The temperature was
gates were still present. Similar state of dispersion character- stable within 0.2 °C over the range used in this study. Strain
izes materials prepared with HD-CNF and LD-CNF. An sweeps were previously performed to investigate the linear
example is reported in Fig. 1. The fiber length distribution viscoelastic response region, and time sweeps were also
was affected by the mixing process, and a reduction of the performed to test the thermal stability of the samples. All
fiber aspect ratio was observed for both CNFs even though tests were carried out in N2 atmosphere.
no quantitative assessment was carried out. The steady-state shear rheology was investigated at 210 °C
The preparation of the nanoreinforced materials could using the rotational rheometer (RMS800) in the low shear rate
affect the matrix properties by inducing some thermome- region and a capillary rheometer (Göttfert Rheograph 2002) in
the high shear rate region. Step rate experiments were
performed under a N2 atmosphere with the cone-plate
Table 3 Compositions of HD-based nanocomposites
geometry (cone angle 0.1 rad) in the shear rate range [0.02,
Unit Fiber concentration 20 s−1]. The temperature was stable within 0.2 °C over the
range used for the rotational rheometer tests. The measure-
wt% 0.50 1.00 3.00 6.00 10.0
ments with the capillary rheometer were performed in the
vol% 0.27 0.53 1.62 3.29 5.58
shear rate range [1, 5000 s−1] and corrected for non-
428 Rheol Acta (2008) 47:425–433

Table 4 Compositions of LD-based nanocomposites

Unit Fiber concentration

wt% 0.25 0.50 1.00 1.50 2.00 3.00 4.50 6.00 10.0
vol% 0.13 0.27 0.53 0.80 1.07 1.62 2.45 3.29 5.58

Newtonian effects (Rabinowitsch correction). A round hole superimpose. It can be noted that G’ increases more than
capillary was used with L/D=30 (diameter=1 mm, length= G” in this concentration range. At low concentrations, the
30 mm, entry angle=180°). This L/D value was large enough storage modulus enhancement is usually attributed to
to allow to neglect the Bagley correction. The rheometer was stiffness imparted by the solid particles that allow efficient
equipped with a pressure transducer able to detect pressures stress transfer, which is mainly controlled by the matrix/
up to 1.4×108 Pa. fiber–matrix interface. At larger loadings (6 and 10 wt%),
the moduli increase steeply. Thus, above a threshold value
(in this case above 3 wt%), a pseudo-solid-like behavior is
Experimental results observed at low frequencies, i.e., G’ and G” data show the
appearance of a plateau, suggesting the formation of some
Linear viscoelasticity interconnected nanofiber network within the matrix. It is
well known from the literature that the interconnected
In the molten state, a strain sweep was performed to structures of anisometric fillers result in an apparent yield
determine the limits of linear viscoelastic regime. At an stress which corresponds, in dynamic measurements, to the
angular frequency 5=6.28 rad/s and at a temperature T= plateau of G’ or G” at low frequencies (Shenoy 1999;
210 °C, the pure reference matrix showed a linear response Utracki 1986; Dealy and Wissbrun 1999). This effect is
up to 100% strain, which was the largest applied strain. more pronounced in G’ than in G”. This is in accordance
Figure 2 shows a reduction of the linear range for the with theoretical expectations and experimental observations
nanoreinforced material with HD fibers starting from 3 wt%, for fiber-reinforced composites (Shaffer et al. 1998;
while for the LD samples starting at 1.5 wt%. Pötschke et al. 2002; Lozano et al. 2004). Indeed, as the
nanofiber content increases, nanofiber–nanofiber interac-
Frequency sweep test—HD carbon nanofibers tions begin to dominate, eventually leading to percolation
and to the formation of some interconnected structure.
Frequency sweep tests at T=210 °C were performed with The magnitudes of the complex viscosity, η*, of the
angular frequencies ranging from 0.1 to 100 rad/s. The reference matrix and composites with different HD-CNF
presence of HD-CNFs produces a small increase in moduli content are shown in Fig. 4. Of course, consistently with
up to 3 wt%, where G’ and G” data follow closely the moduli data, at small loadings (<3 wt%), the presence of
unfilled reference matrix data (PP) throughout the frequen- solid particles perturbs the normal flow of polymer, and, as
cy range as shown in Fig. 3. A “hesitation” is found usually observed in dilute suspensions of rigid particles, the
between 0.5 and 1 wt% as the two data sets nearly
10

LD samples
HD samples
1
Strain

0.1

0.01

0.001
1 10

%wt CNF
Fig. 2 The limiting strain for linear viscoelastic response (T=210 °C,
Fig. 1 A SEM image of HP400R +3 wt% of HD-CNFs 5=6.28 rad/s)
Rheol Acta (2008) 47:425–433 429

Fig. 3 Effect of CNFs HD on 5


10
5
10
the linear viscoelastic response a b
(T=210 °C). a Storage modulus; 4
10
b Loss modulus 4
10
3
10

G" (Pa)
G' (Pa)
3
10
2
10
PP PP
HD 0.5% HD 0.5%
1 2
10 HD 1% 10 HD 1%
HD 3% HD 3%
HD 6% HD 6%
0
10 HD 10% HD 10%
1
10
-1 0 1 2 -1 0 1 2
10 10 10 10 10 10 10 10
ω (rad/s) ω (rad/s)

magnitude of the complex viscosity of the filled system tions, the magnitude of the complex viscosity increases with
increases. Up to 3% loading, the norm of the complex respect to the reference matrix preserving the Newtonian
viscosity of HD-CNF samples shows a dependence on the plateau at low frequencies (Fig. 6). Above 2 wt%, η* data do
frequency similar to that of pure matrix, revealing a not show the Newtonian plateau anymore, and yield stress
Newtonian plateau at low frequencies. A simple vertical appears. This evidence confirms the existence of a
shift is observed as the concentration increases as normally threshold concentration value between 2 and 3 wt%, mar-
observed with conventional microcomposite systems, con- king the onset of a solid-like behavior at low frequencies.
sistently with what is observed for G’ and G”. Above 3 wt%, The enhancement of moduli due to the presence of LD-
η* data do not show the Newtonian plateau at low CNF is significantly higher at lower frequencies, and the
frequencies anymore, and yield stress appears. increase is more pronounced in the storage modulus. Thus,
as it could be expected, the system with lower aspect ratio
Frequency sweep test—LD carbon nanofibers (HD-CNFs) shows a larger percolation concentration with
respect to the system with larger aspect ratio (LD-CNFs). A
A much richer behavior is observed in the case of fibers with similar effect of the aspect ratio was found by Kitano et al.
larger aspect ratio, i.e., with the LD-CNF nanocomposites. In (1984) with polyethylene melts filled with glass fibers. The
this case, the frequency sweep measurements are shown in very recent data proposed by Wang et al. (2006) also show
Fig. 5 (for the sake of clarity, only a selection of the same dependence on the fiber aspect ratio.
concentrations is plotted). It is apparent that the increase of The most striking feature, however, is the distinct
both moduli is much larger than that observed with HD- nonmonotonic dependence of the moduli on the filler
CNFs at the same loading (see Fig. 3), and the pseudo-solid- concentration. Indeed, at a fixed frequency, as the loading
like plateau is observed at lower concentrations (it already increases from 0% (pure polymer), both moduli first increase
appears above 2 wt%). At the lower nanofiber concentra- (at 0.5%) then decrease (at 1.5%), and then they increase
again (at 3%). This feature is even more evident in the
4
10 magnitude of complex viscosity data plotted in Fig. 6.
PP Figure 7 shows the variation of the shear storage
HD 0.5% modulus of composites at 25 °C. It is apparent that both
HD 1%
inclusions improve the modulus as the filler concentration
HD 3%
HD 6% increases, but the effect of the LD-CNFs is substantially
HD 10% larger. Again, the difference between the two set of data can
|η*| (Pa s)

• 3
10 be ascribed to the different aspect ratio of the inclusions.
As already observed in linear viscoelastic data taken in
the molten state, an interesting feature of these data is the
presence of a relative maximum at low composition. This
maximum is clearly visible in the case of LD-CNFs, while
it is hardly detectable in the HD-CNF samples.
2
10
0 1 2
10 10 10 Steady-shear tests
ω (rad/s)
Fig. 4 The magnitude of the complex viscosity (T=210 °C) for the Steady-shear tests were performed only on the LD-CNF
HD-CNF nanocomposites nanocomposites in view of their peculiar response in linear
430 Rheol Acta (2008) 47:425–433

Fig. 5 Effect of LD-CNFs on 5


10 10
5

the linear viscoelasticity a b


(T=210 °C). a Storage modulus; 4
10
b Loss modulus 10
4

3
10

G" (Pa)
G' (Pa)
3
10
2
10
PP PP
LD 0.5% LD 0.5%
1 2
10 LD 1.5% 10 LD 1.5%
LD 3% LD 3%
LD 6% LD 6%
0
10 LD 10% LD 10%
1
10
-1 0 1 2 -1 0 1 2
10 10 10 10 10 10 10 10
ω (rad/s) ω (rad/s)

viscoelasticity. The experiments were carried out in a strain- The most peculiar behavior is the nonmonotonic depen-
controlled rotational rheometer at lower shear rates [0.02– dence of the viscosity on the CNF fraction at low concen-
20 s−1] and in a capillary rheometer at higher shear rates trations, which parallels that observed in the linear
[1–5,000 s−1]. The viscosity data obtained with the two viscoelastic functions. At low shear rates, a distinct New-
instruments perfectly overlap at intermediate shear rates. tonian plateau is visible for loading up to 1.5 wt%. Here, the
At low shear rates, the addition of LD-CNF to the zero shear viscosity of nanocomposites first increases then
polymer matrix produces a general increase of the viscosity decreases, thus confirming the linear viscoelastic features.
with respect to that of the matrix. At lower concentrations, a This feature is apparent in Fig. 9, where the zero shear
Newtonian plateau is found at low shear rates, while at viscosity for the LD-CNF systems that show a Newtonian
higher fractions, yield stress appears at low shear rates (see plateau at low shear rates is plotted as a function of
Fig. 8b). Nanocomposites with solid fraction above 2 wt% nanofiber content.
exhibit strong shear thinning throughout the investigated The trend of the viscosity reported in Fig. 9 is very
shear rate range. Again, data suggest the formation of some similar to that observed in systems showing anisotropic
network structure. Shear thinning behavior of the viscosity mesophases, such as lyotropic rodlike polymers showing
can be associated at low shear rates to the breakdown of the nematic liquid crystalline phases (see the Discussion
network structure and at large shear rates to the possible section).
alignment of the fibers along the flow direction. The similarity with such systems appears to be even
At higher shear rates, the effect of the filler on the stronger if the first normal stress difference is considered.
viscosity is reduced in magnitude, and the nanocomposite The first normal stress difference, N1, was measured during
viscosity tends to the matrix viscosity. The significant start up of steady-shear tests with the rotational rheometer
decrease in viscosity at higher shear rates shows that the equipped with cone and plate fixture. Special care during
difficulty of processing nanofiber systems should not be an sample loading was necessary to have reproducible data. A
issue with those under investigation here.
0.90
5
10
PP 0.85 T=25∞C
shear storage modulus (GPa)

LD 0.5%
LD 1% 0.80
LD 1.5%
LD 2% 0.75
4
10 LD 3%
0.70
|η*| (Pa s)

• LD 6%
LD 10%
0.65

3 0.60
10
0.55
PP+LD CNF
0.50 PP+HD CNF
PP
2
10 0.45
-1 0 1 2 0 1 2 3 4 5 6 7 8 9 10 11
10 10 10 10
ω (rad/s) CNF LD (wt %)
Fig. 6 Effect of LD-CNFs on the magnitude of complex viscosity Fig. 7 Shear storage modulus at T=25 °C for LD and HD CNF
(T=210 °C) nanocomposites
Rheol Acta (2008) 47:425–433 431

Fig. 8 Viscosity for the refer- PP PP


ence matrix (PP) and nanocom- LD 0.25% LD 0.25%
posites with different LD-CNF 4
10 LD 0.5% 4
LD 0.5%
LD 1%
10
loadings (T=210 °C). a Viscos- LD 1%
LD 1.5% LD 1.5%
ity versus shear rate; b Viscosity LD 2%

Viscosity (Pa•s)

Viscosity (Pa•s)
LD 2%
LD 3%
versus shear stress 3
10 LD 6%
3
10
LD 3%
LD 6%
LD 10% LD 10%

2 2
10 10

1
a b
1
10 10
-2 -1 0 1 2 3 4 1 2 3 4 5
10 10 10 10 10 10 10 10 10 10 10 10
-1
shear rate (s ) shear stress (Pa)

waiting time after loading was used to eliminate the effects Discussion
of loading history. During this waiting time, the normal
force generated by sample squeezing was monitored up to The viscosity and normal stress data observed with our LD-
full sample relaxation. The highest CNF concentrations led CNF nanocomposites at 210 °C share several peculiarities
to a maximum waiting period of 1 h before testing. of lyotropic nematic polymers. The introduction of filler
Figure 10 shows the N1 transient for a sample with 6 wt% within a polymeric matrix alters the flow properties of the
LD-CNF at two different shear rates (5 and 10 s−1). The system, and usually the processing becomes more difficult.
shear rate is stepped up after the first 100 strain units. For This aspect is easily verified with rheological characteriza-
the sake of comparison, the transients of the unfilled PP at tion, as a distinct increment of viscosity is usually observed.
the same shear rates are also presented in the figure. The This is not observed, however, in the case of solutions of
effect of the filler with respect to the unfilled sample is rodlike polymers showing a nematic liquid crystalline state
apparent. A clear initial overshoot followed by an under- (Doi and Edwards 1986). In that case, in fact, as the rodlike
shoot and long and oscillating N1 transients were observed particle fraction increases, the zero shear viscosity first
for LD-CNF samples above 1 wt%. increases, then, when a critical concentration is reached, a
Figure 11 reports the average steady state value of the distinct decrement of the viscosity is observed. This
first normal stress difference as function of the shear rate. evidence is related to the transition from an isotropic state
Normal stress signals become measurable only at high to the liquid crystalline phase characterized by a significant
shear rates. At larger LD-CNF composition (6 and 10 wt%), degree of orientational order. Indeed, in the liquid crystal-
N1 attains negative steady-state values. line state, the rodlike molecules are more free to diffuse as
Both long oscillating transient and negative first normal the overall alignment reduces the effects of entanglements
stress difference are typical features of lyotropic rodlike thus increasing the rotational diffusivity. As the concentra-
polymers in the nematic phase. tion of rodlike molecules further increases, the zero-shear

1500
PP 5s-1
PP 10s-1
LD6% 5s-1
LD6% 10s-1

1000
N1 (Pa)

500

0
0 50 100 150 200 250
Strain
Fig. 9 The plateau viscosity at low shear rates versus the LD-CNF Fig. 10 Transients of the first normal stress difference for unfilled PP
composition (T=210 °C) and for LD-CNF 6 wt% plotted versus strain (T=210 °C)
432 Rheol Acta (2008) 47:425–433

3500 PP tropic as well as thermotropic liquid crystalline polymers


LD 0.5%
3000
(e.g., Quijada-Garrido et al. 2000 and references therein).
LD 1%
LD 1.5% All these nonlinear features were here observed for the
2500 LD 2% case of a large aspect ratio CNF system, thus suggesting the
LD 3%
2000 LD 6%
possible formation of a mesophase. Although intriguing for
its processing implications, this parallel should be carefully
N1 (Pa)

LD 10%
1500
examined in view of an important difference between the
1000 two systems. Indeed, it seems now undisputable that the
nanofibers are non-Brownian objects (e.g., Wang et al.
500
2006). On the contrary, rodlike polymers are Brownian, and
0 above a critical concentration under quiescent conditions,
they spontaneously form an ordered nematic phase. No
-500
10
-1 0
10 10
1 evidence is at the moment available on the possible
-1
Shear rate (s ) formation of nanofiber-ordered phases at equilibrium. It
Fig. 11 First normal stress difference measured for LD-CNF nano- should be mentioned, however, that under flow condition,
composites subjected to steady shear flows (T=210 °C) nanofibers can behave as pseudo-Brownian objects, rota-
tional diffusion being caused by casual collisions (Folgar
and Tucker 1984).
viscosity eventually starts to increase again as the ordering A final comment is in order. We did not observe any
beneficial effects cease the way to the progressive increase peculiar feature with the HD-CNF sample. Furthermore,
of viscous drag due to the relative motion between solvent Wang et al. (2006) do not mention any of these phenomena
and rods. A viscosity decrease with increasing fiber fraction for both polystyrene-CNF nanocomposites they investigat-
was already seen in the PET/CNF nanocomposites studied ed. These three systems (our HD-CNF, and the systems
by Ma et al. (2003), but no explanation was suggested for investigated by Wang et al.) are characterized by a lower
the peculiar behavior. Similar features have been observed fiber aspect ratio with respect to that of our LD-CNF
in a system containing multiwalled carbon nanotubes in system. In our opinion, “nematic-like” response can be
aqueous solvent (Song et al. 2003; Song and Windle 2005) observed if the fiber aspect ratio is large enough to promote
and in single-wall nanotubes in superacid (Davis et al. the ordering effects at relatively low compositions. If this is
2004). In these papers, the fact was explained by not the case, the possible formation of a network structure
considering that the phase behavior of carbon nanotube hinders the manifestation of orientational ordering.
systems show many parallels with that of lyotropic
nematogenic rodlike polymer solutions.
Another peculiarity of liquid crystalline polymers is the Final comments
appearance of negative first normal stress difference under
shear at intermediate shear rates (e.g., Kiss and Porter In this paper, we have studied the rheological behavior of
1978). Such a behavior is rather uncommon and is now two carbon nanofiber nanocomposites. The systems differ
well understood (Marrucci and Maffettone 1989; Larson only for the fiber aspect ratio, being otherwise equivalent.
1990). Its origin lays in the disordering effect of shearing The nanofibers are dispersed in a polymer matrix via melt
flow onto the orientational order of a nematic phase. compounding. Linear viscoelasticity shows the formation
Negative first normal stress differences have been measured of a network structure for both systems for compositions
also on single-wall nanotubes in superacid (Davis et al. above a critical value. The nanocomposite characterized by
2004), thus corroborating the LCP similarity of that CNF with larger aspect ratios shows a more pronounced
material on attractive emulsions (Montesi et al. 2004), and increase of moduli and a lower critical composition.
on nanotube suspensions (Lin-Gibson et al. 2004). In the Nonlinear rheology showed that the system with larger
two latter cases, however, the occurrence of negative nanofiber aspect ratio has several peculiarities similar to
normal stresses was related to vorticity alignment of polymers in the nematic phase. This aspect can have
macroscopic aggregates. interesting practical implications in the applications, as
Finally, as far as the transient behavior of first normal ordered state is usually characterized by exceptional
stress difference is concerned, the evidence of long mechanical properties.
oscillating transients is again very similar to the response
of ordered mesophases. Indeed, oscillations in N1 have
been found for lyotropic nematic solutions of poly(benzyl Acknowledgements This work was supported by FIRB MAPIO
glutamate) (PBG), hydroxypropylcellulose, and other lyo- NANO.
Rheol Acta (2008) 47:425–433 433

References Lozano K, Yang S, Zeng Q (2004) Rheological analysis of vapor-


grown carbon nanofiber-reinforced polyethylene composites. J
Appl Polym Sci 93:155–162
Choi YK, Sugimoto KI, Song SM, Endo M (2005) Mechanical and Ma HM, Zeng JJ, Realff ML, Kumar S, Schiraldi DA (2003)
thermal properties of vapor-grown carbon nanofiber and poly- Processing, structure, and properties of fibers from polyester/
carbonate composite sheets. Mater Lett 59(27):3514–3520 carbon nanofiber composites. Compos Sci Technol 63(11):1617–
Davis VA, Ericson LM, Parra-Vasquez ANG, Fan H, Wang YH, Prieto 1628
V, Longoria JA, Ramesh S, Saini RK, Kittrell C, Billups WE, Marrucci G, Maffettone PL (1989) Description of the liquid crystalline
Adams WW, Hauge RH, Smalley RE, Pasquali M (2004) Phase phase of rodlike polymers at high shear rates. Macromolecules
behavior and rheology of SWNTs in superacids. Macromolecules 22:4076–4082
37(1):154–160 Montesi A, Pena AA Vorticity alignment and negative normal stresses
Dealy M, Wissbrun KF (1999) Melt rheology and its role in plastic in sheared attractive emulsions. Phys Rev Lett 92(5):058303
processing theory and application. Kluwer, Dordrecht Pötschke P, Fornes TD, Paul DR (2002) Rheological behavior of
Doi M, Edwards SF (1986) The theory of polymer dynamics. multiwalled carbon nanotube/polycarbonate composites. Polymer
Clarendon, Oxford 43:3247–3255
Folgar F, Tucker CL (1984) Orientation behaviour of fibres in Quijada-Garrido I, Siebert H, Friedrich C, Schmidt C (2000) Flow
concentrated suspensions. J Reinf Plast Compos 3:98–119 behavior of two side-chain liquid crystal polymers studied by
Gao Y, He P, Lian J, Wang LM, Qian D, Zhao J, Wang W, Schulz MJ, transient rheology. Macromolecules 33(10):3844–3854
Zhang J, Zhou XP, Shi DL (2006) Improving the mechanical Sandler J, Windle AH, Werner P, Altstadt V, Es MV, Shaffer MSP
properties of polycarbonate nanocomposites with plasma-modified (2003) Carbon-nanofibre-reinforced poly(ether ether ketone)
carbon nanofibers. J Macromol Sci, Phys 45(4):671–679 fibres. J Mater Sci 38(10):2135–2141
Gauthier C, Chazeau L, Prasse T, Cavaille JY (2005) Reinforcement Shaffer MSP, Fan X, Windle AH (1998) Dispersion and packing of
effects of vapour grown carbon nanofibres as fillers in rubbery carbon nanotubes. Carbon 36(11):1603–1612
matrices. Compos Sci Technol 65(2):335–343 Shen J, Han XM, Lee LJ (2006) Nanoscaled reinforcement of
Kang IP, Heung YY, Kim JH, Lee JW, Gollapudi R, Subramaniam S, polystyrene foams using carbon nanofibers. J Cell Plast 42
Narasimhadevara S, Hurd D, Kirikera GR, Shanov V, Schulz MJ, (2):105–126
Shi DL, Boerio J, Mall S, Ruggles-Wren M (2006) Introduction Shenoy AV (1999) Rheology of filled polymer systems. Kluwer,
to carbon nanotube and nanofiber smart materials. Compos Eng Dordrecht
37(6):382–394 Song W, Windle AH (2005) Isotropic–nematic phase transition of
Kelarakis A, Yoon K, Somani R, Sics I, Chen XM, Hsiao BS, Chu B dispersions of multiwall carbon nanotubes. Macromolecules
(2006) Relationship between structure and dynamic mechanical 38:6181–6188
properties of a carbon nanofiber reinforced elastomeric nano- Song W, Kinloch I, Windle AH (2003) Nematic liquid crystallinity of
composite. Polymer 47(19):6797–6807 multiwall carbon nanotubes. Science 302:1363–1363
Kiss G, Porter RS (1978) Rheology of concentrated solutions of poly Utracki LA (1986) Flow and flow orientation of composites contain-
(g -benzyl-glutamate). J Polym Sci, Polymer Symposia 65:193– ing anisometric particles. Polym Compos 7(5):274–282
211 Wang YR, Xu JH, Bechtel SE, Koelling KW (2006) Melt shear
Kitano T, Kataoka T, Nagatsuka Y (1984) Shear flow rheological rheology of carbon nanofiber/polystyrene composites. Rheol
properties of vinylon- and glass fiber -reinforced polyethylene Acta 45(6):919–941
melts. Rheol Acta 23(1):20–23 Xu J, Donohoe JP, Pittman CU (2004) Preparation, electrical and
Kumar S, Dang TD, Arnold FE, Bhattacharyya AR, Min BG, Zhang mechanical properties of vapor grown carbon fiber (VGCF)/vinyl
XF, Vaia RA, Park C, Adams WW, Hauge RH, Smalley RE, ester composites. Composites Part A 35(6):693–701
Ramesh S, Willis P (2002) Synthesis, structure, and properties of Xu JH, Chatterjee S, Koelling KW, Wang YR, Bechtel SE (2005)
PBO/SWNT composites. Macromolecules 35(24):9039–9043 Shear and extensional rheology of carbon nanofiber suspensions.
Larson RG (1990) Arrested tumbling in shearing flows of liquid Rheol Acta 44(6):537–562
crystal polymers. Macromolecules 23:3983–3992 Yang S, Taha-Tijerina J, Serrato-Diaz V, Hernandez K, Lozano K
Lin-Gibson S, Pathak JA, Grulke EA, Wang H, Hobbie EK (2004) (2007) Dynamic mechanical and thermal analysis of aligned
Elastic flow instability in nanotube suspensions. Phys Rev Lett vapor grown carbon nanofiber reinforced polyethylene. Compo-
92(4):048302 sites Part B 38(2):228–235

You might also like